Skip to main content
Log in

Most Probable Paths for Anisotropic Brownian Motions on Manifolds

  • Published:
Foundations of Computational Mathematics Aims and scope Submit manuscript

Abstract

Brownian motion on manifolds with non-trivial diffusion coefficient can be constructed by stochastic development of Euclidean Brownian motions using the fiber bundle of linear frames. We provide a comprehensive study of paths for such processes that are most probable in the sense of Onsager–Machlup, however with path probability measured on the driving Euclidean processes. We obtain both a full characterization of the resulting family of most probable paths, reduced equation systems for the path dynamics where the effect of curvature is directly identifiable, and explicit equations in special cases, including constant curvature surfaces where the coupling between curvature and covariance can be explicitly identified in the dynamics. We show how the resulting systems can be integrated numerically and use this to provide examples of most probable paths on different geometries and new algorithms for estimation of mean and infinitesimal covariance.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

Notes

  1. https://bitbucket.org/stefansommer/jaxgeometry.

References

  1. J. Bradbury, R. Frostig, P. Hawkins, M. J. Johnson, C. Leary, D. Maclaurin, G. Necula, A. Paszke, J. VanderPlas, S. Wanderman-Milne, and Q. Zhang. JAX: Composable transformations of Python+NumPy programs, 2018.

  2. K. Engø, A. Marthinsen, and H. Z. Munthe-Kaas. The diffman package on github. https://github.com/kenthe/DiffMan. Accessed: 2021-10-19.

  3. M. Frechet. Les éléments aléatoires de nature quelconque dans un espace distancie. Ann. Inst. H. Poincaré, 10:215–310, 1948.

    MathSciNet  Google Scholar 

  4. T. Fujita and S.-i. Kotani. The Onsager-Machlup function for diffusion processes. Journal of Mathematics of Kyoto University, 22(1):115–130, 1982.

    MathSciNet  Google Scholar 

  5. P. Hansen, B. Eltzner, S. F. Huckemann, and S. Sommer. Diffusion Means in Geometric Spaces. arXiv:2105.12061, May 2021.

  6. P. Hansen, B. Eltzner, and S. Sommer. Diffusion Means and Heat Kernel on Manifolds. Geometric Science of Information 2021, Feb. 2021.

  7. L. Hörmander. Hypoelliptic second order differential equations. Acta Math., 119:147–171, 1967.

    Article  MathSciNet  Google Scholar 

  8. E. P. Hsu. Stochastic analysis on manifolds, volume 38 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2002.

  9. P. Malliavin. Stochastic calculus of variation and hypoelliptic operators. In Proceedings, International Symposium on SDE, Kyoto, 1976.

  10. R. Montgomery. A tour of subriemannian geometries, their geodesics and applications, volume 91 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2002.

  11. X. Pennec. Intrinsic Statistics on Riemannian Manifolds: Basic Tools for Geometric Measurements. J. Math. Imaging Vis., 25(1):127–154, 2006.

    Article  MathSciNet  Google Scholar 

  12. R. W. Sharpe. Differential geometry, volume 166 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1997. Cartan’s generalization of Klein’s Erlangen program, With a foreword by S. S. Chern.

  13. I. Shigekawa. On stochastic horizontal lifts. Z. Wahrsch. Verw. Gebiete, 59(2):211–221, 1982.

    Article  MathSciNet  Google Scholar 

  14. S. Sommer. Anisotropic Distributions on Manifolds: Template Estimation and Most Probable Paths. In Information Processing in Medical Imaging, volume 9123 of Lecture Notes in Computer Science, pages 193–204. Springer, 2015.

  15. S. Sommer. Evolution Equations with Anisotropic Distributions and Diffusion PCA. In F. Nielsen and F. Barbaresco, editors, Geometric Science of Information, number 9389 in Lecture Notes in Computer Science, pages 3–11. Springer International Publishing, 2015.

  16. S. Sommer. Anisotropically Weighted and Nonholonomically Constrained Evolutions on Manifolds. Entropy, 18(12):425, 2016.

    Article  ADS  MathSciNet  Google Scholar 

  17. S. Sommer. An Infinitesimal Probabilistic Model for Principal Component Analysis of Manifold Valued Data. Sankhya A, Aug. 2018.

  18. S. Sommer and A. M. Svane. Modelling anisotropic covariance using stochastic development and sub-Riemannian frame bundle geometry. Journal of Geometric Mechanics, 9(3):391–410, 2017.

    Article  MathSciNet  Google Scholar 

  19. D. W. Stroock and S. R. S. Varadhan. On the support of diffusion processes with applications to the strong maximum principle. In Proceedings of the Sixth Berkeley Symposium on Mathematical Statistics and Probability (Univ. California, Berkeley, Calif., 1970/1971), Vol. III: Probability theory, pages 333–359. Univ. California Press, Berkeley, Calif., 1972.

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Stefan Sommer.

Additional information

Communicated by Carola-Bibiane Schönlieb.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Erlend Grong is supported by the grant GeoProCo from the Trond Mohn Foundation Grant TMS2021STG02 (GeoProCo). Stefan Sommer is supported by the Villum Foundation Grants 22924 and 40582, and the Novo Nordisk Foundation Grant NNF18OC0052000.

Appendices

Definition of Sub-Riemannian Geometry

We give a quick introduction to sub-Riemannian geometry and refer to [10] for details. A sub-Riemannian manifold is a triple \((M, E,\rho )\) where M is a connected manifold, E is a subbundle of the tangent bundle TM and \(\rho = \langle \cdot , \cdot \rangle _\rho \) is a metric tensor defined only on E. This tensor defines a vector bundle morphism \(\sharp ^\rho : T^*M \rightarrow E \subseteq TM\) given by

$$\begin{aligned} \alpha (v) = \langle \sharp ^\rho \alpha , v \rangle _\rho . \end{aligned}$$

Consequently, we obtain a positive semi-definite symmetric tensor \(\rho ^* = \langle \cdot , \cdot \rangle _{\rho ^*}\) on \(T^*M\) defined by

$$\begin{aligned} \langle \alpha , \beta \rangle _{\rho ^*} = \langle \sharp ^\rho \alpha , \sharp ^\rho \beta \rangle _\rho . \end{aligned}$$

This tensor degenerates along the subbundles \({{\,\mathrm{Ann}\,}}(E) \subseteq T^*M\) of covectors vanishing on E. It follows that a sub-Riemannian manifold can equivalently be defined as a connected manifold with positive, semi-definite cometric \(\rho ^*\) that degenerates along a subbundle.

An absolutely continuous curve \(\gamma :[0,T] \rightarrow M\) is called horizontal if \({{\dot{\gamma }}}(t) \in E_{\gamma (t)}\) for almost every t. For such a curve, we define its length to be

$$\begin{aligned} L^\rho (\gamma ) = \int _0^T | {{\dot{\gamma }}}|_\rho (t)\, {d}t. \end{aligned}$$

This length is invariant under reparametrization, so we can restrict our considerations to the case \(T =1\).

For any \(x, y \in M\), we define

$$\begin{aligned} d_\rho (x,y) = \inf \left\{ L^\rho (\gamma ) \, : \, \begin{array}{c} \gamma :[0,1] \rightarrow M \text { horizontal} \\ \gamma (0) = x, \gamma (1) = y \end{array} \right\} . \end{aligned}$$

We notice that if there are no horizontal curves connecting the two points, then \(d_\rho (x,y) =\infty \). Let \(x \in M\) be a given point. We define \({\mathcal {C}}(x)\) as the space of all horizontal curves defined on [0, 1], with \(L^2\)-derivative, that start in x. This collection has a natural structure of a Hilbert manifold, see [10, Chapter 5.1] for details. Define a mapping

$$\begin{aligned} \varPi : {\mathcal {C}}(x) \rightarrow M, \qquad \gamma \mapsto \gamma (1). \end{aligned}$$

Define \({\mathcal {C}}(x,y) = \varPi ^{-1}(y)\). A point \(\gamma \in {\mathcal {C}}(x,y)\) is called regular if \(\varPi _{*,\gamma }:T_\gamma {\mathcal {C}}(x) \rightarrow T_y M\) is surjective. Otherwise, \(\gamma \) is called singular or abnormal curves.

Assume that \({\mathcal {C}}(x,y)\) non-empty. Define \(F: {\mathcal {C}}(x,y) \rightarrow {\mathbb {R}}\) by \(\gamma \mapsto L^\rho (\gamma )\). We look at minimal elements in \({\mathcal {C}}(x,y)\) with respect to F. If \(\gamma \) is a regular curve, then \({\mathcal {C}}(x,y)\) locally has the structure of a Hilbert manifold around \(\gamma \) by the inverse function theorem. Hence, any regular minimal element must be a critical, i.e., we must have \(F_{*,\gamma } = 0\). Such curves are called normal geodesics, and will always be locally length minimizing. It can be shown that all such curves, up to reparametrization, be found as a projecting of a solution of a Hamiltonian system. The Hamiltonian is given by

$$\begin{aligned} P(\alpha ) = \frac{1}{2} \langle \alpha , \alpha \rangle _{\rho ^*}, \qquad \alpha \in T^*M. \end{aligned}$$

In conclusion, length minimizers are either normal geodesics or abnormal curves. These classes of curves are not necessarily disjoint.

We say that E is bracket-generating if for every point \(x \in M\),

$$\begin{aligned} {{\,\mathrm{span}\,}}\{ X_i, [X_i,X_j], [X_i, [X_j, X_k]], \dots , \} |_x \in T_xM, \qquad X_i \in \varGamma (E), \end{aligned}$$

that is, if sections of E generate the entire tangent bundle TM. If this condition holds, then any pair of points can be connected by a horizontal curve. The value of \(d_\rho \) is always finite, and furthermore, it induces the same topology as the manifold topology.

Remark 7

Let L be a second order operator on M without constant term, such that for any pair of smooth functions \(f,g \in C^\infty (M)\),

$$\begin{aligned} L(fg) - fLg - gL f = \langle df, dg \rangle _{\rho ^*}. \end{aligned}$$

In other words, locally, L can always be written as \(L = \sum _{j=1}^{{{\,\mathrm{rank}\,}}E} V_i^2 + V_0\), where \(V_1, \dots , V_{{{\,\mathrm{rank}\,}}E}\) is a local orthonormal basis of \((E,\rho )\). If E is bracket generating, then L is hypoelliptic [7] and its heat semigroup \(p_t(x;y)\) has a strictly positive density [19].

Sub-Riemannian Normal Geodesics on \(({{\,\mathrm{Sym}\,}}^+ TM, E, \rho )\)

By the discussion in Sect. 4.2, it follows that we can write a normal geodesic as \(\varSigma (t) = f(t)^{-1} S^2 f(t)\) where f(t) is a normal geodesic in \(({{\,\mathrm{O}\,}}(TM), {\mathcal {H}}, \rho _S)\). We do the computations here.

Recall the definition of the vector fields \(H_a\), \(a \in {\mathbb {R}}^n\) and \(\xi _A\), \(A \in {{\,\mathrm{\mathfrak {so}}\,}}(n)\) in Sect. 3. We introduce corresponding Hamiltonian functions

$$\begin{aligned} P_a(\alpha ) = \alpha (H_a), \qquad Q_A(\alpha ) = \alpha (\xi _A), \qquad \alpha \in T^* {{\,\mathrm{O}\,}}(TM). \end{aligned}$$

Our formulas in (3.3), then give corresponding relations in terms of Poisson brackets

$$\begin{aligned} \{ P_a, P_b\} = Q_{{\underline{R}}(a,b)}, \qquad \{ Q_A, P_a\} = - P_{Aa}, \qquad \{ Q_A, Q_B\} = - Q_{[A,B]}. \end{aligned}$$

Since \(H_{Se_1}, \dots , H_{Se_d}\) is a global orthonormal basis, we have that the sub-Riemannian Hamiltonian is given by

$$\begin{aligned} P = \frac{1}{2} \sum _{j=1}^d P_{Se_j}^2. \end{aligned}$$

Let \(\lambda (t) = e^{t\mathbf {P}}(\lambda _0)\) be a solution in \(T^* FM\) along f(t) in FM and define curves v(t) in \({\mathbb {R}}^d\) and A(t) in \({{\,\mathrm{\mathfrak {so}}\,}}(d)\) by

$$\begin{aligned} P_a(t) = P_a(\lambda (t)) = \langle S^{-2} v(t), a \rangle , \qquad Q_B(t) = Q_B(\lambda (t)) = - \langle A(t), B \rangle . \end{aligned}$$

Then along a solution, we have

$$\begin{aligned} \langle S^{-2} \dot{v}, a \rangle&= \dot{P}_a = \{ P_a, P\} = \sum _{j=1}^d P_{Se_j} Q_{{\underline{R}}(a, Se_j)}\\&= - \sum _{j=1}^d \langle S^{-2} v, Se_j \rangle \langle A, {\underline{R}}(a, Se_j) \rangle \\&= \sum _{j=1}^d \langle S^{-1} v, e_j \rangle \langle {\underline{R}}(A)S e_j, a\rangle = \langle {\underline{R}}(A) v, a\rangle \\ - \langle \dot{A}(t) , B \rangle&= - \dot{Q}_B \\&= - \{ Q_B, P \} = \sum _{j=1}^d P_{Se_j} P_{BSe_j} \\&= \sum _{j=1}^d \langle S^{-2}v, Se_j \rangle \langle BS e_j ,S^{-2} v \rangle \\&= \langle B v, S^{-2}v \rangle \\&= \langle v \wedge S^{-2} v, B \rangle . \end{aligned}$$

In summary, \(\dot{v} = S^2 {\underline{R}}(A) v\) and \(\dot{A} = v \wedge S^{-2} v\).

Let \(\beta \in \varGamma (T^* {{\,\mathrm{O}\,}}(TM))\) be a one-form on \({{\,\mathrm{O}\,}}(TM)\) and define the corresponding vertical lift \({{\,\mathrm{vl}\,}}\beta \in \varGamma (T(T^* {{\,\mathrm{O}\,}}(TM)))\) by

$$\begin{aligned} {{\,\mathrm{vl}\,}}\beta |_\alpha = \frac{{d}}{{d}t} (\alpha + t\beta _f) |_{t=0}, \qquad \alpha \in T_f^* {{\,\mathrm{O}\,}}(TM). \end{aligned}$$

Let \(\vartheta \) be the Liouville one form \(\vartheta |_\alpha = \pi ^* \alpha \) with canonical symplectic form \(\sigma = - d\vartheta \). Observe that \(\sigma ({{\,\mathrm{vl}\,}}\beta , \, \cdot \, ) = - (\pi ^* \beta )(\, \cdot \,)\). Then

$$\begin{aligned} dP({{\,\mathrm{vl}\,}}\beta )&= \sum _{j=1}^d P_{S e_j} \beta (H_{S e_j} ) = \sigma ( \mathbf {P}, {{\,\mathrm{vl}\,}}\beta ) = \beta (\pi _* \mathbf {P}). \end{aligned}$$

If we consider this along the curve, we have

$$\begin{aligned} \sum _{j=1}^d P_{S e_j} \beta (H_{S e_j} ) |_{\lambda (t)}&= \sum _{j=1}^d \langle S^{-1}v(t), e_j \beta (H_{S e_j} ) |_{f(t)} = \beta (H_{v(t)} ) |_{f(t)} \\&= \beta (\pi _* \mathbf {P}) |_{\lambda (t)} = \beta (\dot{f}(t)) \end{aligned}$$

It follows that

$$\begin{aligned}&\dot{f}(t) = H_{v(t)}, \\&\dot{v}(t) = S^2 {\underline{R}}(f(t))(A(t)) v(t), \\&\dot{A}(t) =v(t) \wedge S^{-2} v(t). \end{aligned}$$

We see that these are exactly the equations of found in the proof of Theorem 1 without the condition \(A(T) = 0\).

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Grong, E., Sommer, S. Most Probable Paths for Anisotropic Brownian Motions on Manifolds. Found Comput Math 24, 313–345 (2024). https://doi.org/10.1007/s10208-022-09594-4

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10208-022-09594-4

Keywords

Mathematics Subject Classification

Navigation